Sn Graphane Alloys

Nov 12, 2014 - In these two-dimensional alloys, the germanium atom is terminated ... the Ge and Sn atoms in these graphane alloys rapidly oxidize upon...
0 downloads 0 Views 2MB Size
Subscriber access provided by West Virginia University | Libraries

Article

Synthesis and Stability of Two-dimensional Ge/Sn Graphane Alloys Maxx Q. Arguilla, Shishi Jiang, Basant Chitara, and Joshua E. Goldberger Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/cm502755q • Publication Date (Web): 12 Nov 2014 Downloaded from http://pubs.acs.org on November 19, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 13

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Synthesis and Stability of Two-dimensional Ge/Sn Graphane Alloys Maxx Q. Arguilla†, Shishi Jiang†, Basant Chitara†, Joshua E. Goldberger†. †

Department of Chemistry and Biochemistry, The Ohio State University, Columbus, Ohio 43210-1340

ABSTRACT: There has been considerable interest in the germanium and tin graphane analogues due to their potential as optoelectronic building blocks, and novel topological materials. Here, we have synthesized for the first time alloyed germanium/tin graphane analogues from the topochemical deintercalation of CaGe2-2xSn2x (x = 0-0.09) in aqueous HCl. In these two-dimensional alloys, the germanium atom is terminated with hydrogen while tin is terminated with hydroxide. With greater tin incorporation, the band gap systematically shifts from 1.59 eV in GeH down to 1.38 eV for Ge0.91Sn0.09H0.91(OH)0.09, which allows for more sensitive photodetection at lower energies. In contrast to germanane’s oxidation-resistance, the Ge and Sn atoms in these graphane alloys rapidly oxidize upon exposure to air. This work demonstrates the possibility of creating functional tin-incorporated group IV graphane analogues.

Introduction Layered two-dimensional (2D) materials have shown considerable potential as novel building blocks for semiconductor device applications including next generation electronics,1 photodetectors,2 solar cells,3 light emitting diodes,4 and photocatalysts.5 To these ends, the group IV graphane analogues, in which every atom exists in a ligand-terminated puckered hexagonal honeycomb geometry, have attracted recent interest as 2D materials whose properties can be tuned via covalent chemistry of the surface-ligand.6 For example, depending on the nature of the surface functionalization, the 2D silicanes feature band gaps ranging from 2.5-3.1 eV,7 the germananes have ~1.6-1.8 eV gaps.6a,6c,6e Furthermore, recent theoretical calculations have predicted the 2D stannananes to be quantum spin hall topological insulators with 0.3-0.4 eV gaps.8 Despite these exciting predictions there have been no experimental reports of any tincontaining graphane analogues. It is unclear if many of the proposed ligand terminations on tin in a 2D honeycomb lattice are even chemically stable. Indeed, previous studies on 2D silicane/germanane alloys have shown that silicon has a strong chemical preference for hydroxide termination, whereas germanium is preferentially terminated by hydrogen.9 The creation of 2D germanium/tin graphane alloys would allow a better understanding of how the electronic structure, optical properties, and the degree of spin orbit coupling can be tuned to realize enhanced optoelectronic properties and novel topological phenomena. The creation of three-dimensional germanium/tin semiconductor alloys has recently gained a lot of research traction, and has shown the feasibility of alloying germanium and tin together to optimize properties. For example, theoretical10 and experimental results10d,11 show that the incorporation of 6-8% tin into a germanium lattice can change the nominally 0.67 eV indirect gap of Ge to a

0.5-0.6 eV direct gap, thus leading to significantly enhanced light emission.11d,12 Consequently germanium/tin alloys allow have attracted considerable interest as active materials for wavelength tunable infrared photodetectors11b,13 and light-emitting devices.14 Just as alloying germanium and tin has proven to be a useful strategy for tuning the electronic structure in these 3D semiconductors, similar phenomena should be expected in a 2D lattice. The chemical route to prepare these Group IV graphane analogues relies on the topotactic deintercalation of alloys of precursor layered Zintl phases. For example, we have previous established that CaGe2 can be topochemical transformed in acidic solutions to produce hydrogen-terminated germanane (GeH).6a 6e Preparation of these alloyed 2D Group IV graphanes requires the synthesis of similar alloyed Zintl phases. There are a couple reports of the synthesis of CaGe2-2xSn2x (x=0.2-0.5)15 however, these Zintl phases often have a nonhomongeneous alloy distribution. Furthermore, no such study on the topochemical deintercalation of these materials has been performed. Herein, we demonstrate for the first time the creation of 2D Ge1-xSnxH1-x(OH)x (x=0-0.09) graphane analogues, in multilayered van der Waals crystals. This represents the maximum quantity of tin that could be incorporated in the starting CaGe2=2xSn2x Zintl phase as a homogeneously distributed alloy. After deintercalation in aqueous HCl, we show that Sn is terminated with –OH, while the Ge remains terminated with –H. The band gap of the resulting 2D material systematically shifts from 1.58 eV in germanane (GeH) down to 1.38 eV with 9% tin. This change in band gap enables photodetection at energies between 1.38-1.58 eV for the 9% tin sample. These germanium/tin graphane alloys start oxidizing after exposure to air, as evidenced by the shift in the absorption spectrum and XPS. Taken together, this work shows that the topotactic deintercalation of Zintl phase precur-

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

sors is a viable route for creating band-gap tunable germanium/tin graphane alloys.

Experimental Details Synthesis of CaGe2-2xSn2x. In a typical reaction, Ca, Ge and Sn were loaded in stoichiometric amounts into a quartz tube and evacuated on a Schlenk line to milliTorr pressures. The quartz tube was sealed under vacuum using a hydrogen-oxygen torch, annealed at 950 to 1050 o C for 16 to 20 hours, and cooled to room temperature over 1 to 5 days. The preparation and handling of all reagents were all done in an Ar-filled glovebox. Germanium (Ge, 99.999%, Acros), calcium (Ca, 99%, Acros) and tin (Sn, STREM Chemicals 99.8%) were purchased and used without further purification. Synthesis of Ge1-xSnxH1-x(OH)x. CaGe2-2xSn2x crystals were stirred in concentrated HCl(aq) for 5 to 14 days at 40 oC. To purify the Ge1-xSnxH1-x(OH)x samples, the resulting product was washed with Milli-Q H2O followed by distilled acetonitrile. Samples were dried at room temperature on a Schlenk line and kept in an Ar-filled glovebox. X-ray Diffraction of CaGe2-2xSn2x and Ge1-xSnxH1x(OH)x. Flat Plane and Capillary mode X-ray diffraction patterns of the samples (germanium was added as an internal standard) were collected on a Bruker D8 powder X-ray diffractometer (Sealed Cu X-ray tube: 40 kV and 50mA). The lattice parameters of the Zintl phases were obtained through Rietveld Analysis of the lab data. Fourier-Transform Infrared (FTIR) Spectroscopy of Ge1-xSnxH1-x(OH)x. Mid-IR (400-4000 cm-1) transmittance spectra of KBr-dispersed samples were collected on a Perkin-Elmer Frontier Dual-Range FIR/Mid-IR spectrometer that was loaded in an Ar-filled glovebox. Raman Spectroscopy of Ge1-xSnxH1-x(OH)x. Raman scattering spectra were collected using a Renishaw InVia Raman equipped with a CCD detector. The Raman spectra were collected using a 785 nm near-IR diode laser as illumination source. Tin Percentage Determination of Ge1-xSnxH1-x(OH)x. The various tin content in Ge1-xSnxH1-x(OH)x were obtained through X-ray fluorescence measurements using a DELTA hand-held X-ray fluorescence analyzer. Germanium and tin powder mixtures with 0%, 1%, 2.5%, 5%, 7.5% and 10% tin were used as calibration standards. The air sensitivity of the Zintl phase precluded the XRF measurements and the determination of the Sn percentage of the Zintl phase precursors. Absorption measurements of Ge1-xSnxH1-x(OH)x. All absorption measurements were collected using a Perkin Elmer Lambda 950 UV/Vis NIR spectrophotometer. Band gap determination was performed using diffuse reflectance absorbance setup that utilizes an integrating sphere. For the oxidation studies, to confirm the changes occurred throughout the bulk of the sample, transmission measurements were performed on ~3µm-thick samples exfoliated onto Polydimethylsiloxane, in addition to diffuse reflectance measurements. The thickness of the exfoliated flakes was calibrated using a KLA-Tencor Alpha-Step profilometer.

Page 2 of 13

X-Ray Photoelectron Spectroscopy. The X-Ray Photoelectron spectra of the samples were taken using a Kratos Axis Ultra X-Ray Photoelectron Spectrometer with a monochromated (Al) X-ray source. Energy calibration was performed using the C 1s peak. Photocurrent Measurements of GeH and Ge0.91Sn0.09H0.91(OH)0.09. Gold (20 nm) and Silver (100 nm) contacts were deposited onto the bulk flakes via Ebeam evaporation through a shadow mask. The wavelength-dependent photocurrent response at 3V was measured using a Keithley 6340 Sub-femtoamp meter and a variable wavelength tungsten filament light source equipped with a monochromator (PV Measurements). The intensity of the light source was calibrated using a Silicon photodiode.

Results and Discussion

Figure 1. (a) Schematic of the topochemical deintercalation of CaGe2-2xSn2x to Ge1-xSnxH1-x(OH)x (Ca: yellow, Ge: blue, H: black, O: red and Sn: green). (b) a/b- and (c) c-lattice parameters of CaGe2-2xSn2x to Ge1-xSnxH1-x(OH)x. (d) Optical image of Ge0.91Sn0.09H0.91(OH)0.09 crystals on a 2.5 mm grid paper. (e) Capillary mode powder XRD pattern of Ge1xSnxH1-x(OH)x (x=0-0.09). The internal Ge standard is labeled with an asterisk ( * ).

The preparation of the 2D germanium/tin graphane alloys was accomplished by the topotactic deintercalation of CaGe2-2xSn2x (x =0, 0.04, 0.07, 0.09) (Fig 1a). First, crystals of CaGe2-2xSn2x (Fig. 1a) were synthesized by

ACS Paragon Plus Environment

Page 3 of 13

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

sealing stoichiometic amounts of Ca, Ge and Sn inside a quartz tube, annealing to 950-1050 oC, and cooling over a period of 3-5 days. The resulting CaGe2-2xSn2x crystals formed two-layer hexagonal P63mc unit cells similar to αCaGe216 where the c-axis corresponds to the out-ofplane direction while the a/b-axes correspond to the inplane lattice directions, as determined via X-ray diffraction (XRD) (Fig. S1a). A systematic increase in both the a/b- (in-plane) and the c-lattice (out-of-plane) parameters both confirm the incorporation of Sn in the Zintl phase (Fig. 1b,c). In addition, the narrow Full Width Half Maximum (FWHM) of the (100) peak (Fig. S1b) around a 2Theta value of 0.1o highlights the homogeneous distribution of tin in the lattice. The crystal structure of these samples was confirmed via Rietveld analysis (Table S16). Rather than homogeneously distributing onto both germanium 2a and 2b sites, Rietveld analysis gave a much better fit when Sn preferentially occupied the Germanium 2b position only, similar to what was reported for the case of CaSn0.5Ge1.5. 15a The only other phase was residual tin impurity, which is subsequently removed during the HCl deintercalation step. Rietveld analysis suggested a ~4-5% phase fraction of tin impurity for all Zintl alloys. After extensive optimization of the synthetic parameters, the maximum amount of tin that could be incorporated into the starting Zintl phase while maintaining this homogeneous distribution was 9%. These CaGe2-2xSn2x crystals were converted into 2D Ge1-xSnxH1-x(OH)x (x=0-0.09) by the topotactic deintercalation in aqueous HCl at -40 oC for 5-14 days (Fig. 1a). In this process, the Ca2+ ions in between the layers are removed via the formation of a water-soluble CaCl2 species while the Ge atoms are terminated by H+ from HCl and the Sn atoms are terminated by OH- from water. After HCl treatment, the product was filtered and washed with water to remove residual CaCl2, and then acetonitrile, and subsequently dried under vacuum. The resulting platelet-shaped crystallites were 2-3 mm in length and width, and 10-100 µm in thickness (Fig. 1d). Calibrated X-ray Fluorescence measurements of these crystals confirm the Ge:Sn ratio (Fig. S2). XRD of these platelets confirmed the transformation into a 2D Ge1-xSnx analog (Fig 1e). Similar to GeH,6a the diffraction can be fit to a 2H unit cell where there are two Ge1-xSnx layers per hexagonal c-axis unit cell d-spacing. In this unit cell, the a/b-axis corresponds to the in-plane direction, parallel to the 2D germanium/tin layer while the c-axis corresponds to the out-of-plane direction. The position of the (002) peak and the (100) peaks were determined from an XRD pattern using an internal germanium standard. Compared to the original CaGe2-2xSn2x unit cell parameters, the a/b- direction is slightly contracted (Fig. 1b) and the c-direction is significantly expanded in 2D Ge1xSnx (Fig. 1c). Furthermore, there is a significant increase in the FWHM of the (002) and the (100) peaks after the deintercalation process, with values close to what was observed in GeH (Fig. S3). Also, after deintercalation there is an increase in all the unit cell parameters with increasing Sn, suggesting that Sn remains in the lattice. The observed ~0.1 Å contraction in the a/bdirection is consistent with the change observed for going from CaGe2 to GeH. However, the 1.2 Å c-axis expansion is much larger than the ~0.4 Å increase that is

observed when replacing a Ca2+ in CaGe2 with 2 Ge-H bonds between each layer.6a This suggests the presence of –OH termination of the Sn substituents which contributes to the larger interlayer c-axis spacing. To further elucidate the identity of the surface terminating ligand, transmission-mode FTIR measurements were performed (Fig. 2a). Upon tin incorporation, a broad O– H stretching mode at ~3450 cm-1 as well as a broad, intense Sn–O stretching mode centered at ~560 cm-1 are observed.17 Sn-H stretching modes centered at ~17001900 cm-1 were not observed at any concentration of Sn.18 Additionally, in all samples, extremely strong Ge– H stretching modes centered at ~2000 cm-1, weaker wagging modes that occur at 570 cm-1, 507 cm-1, and 475 cm-1, and weak vibrational modes at 770 and 825 cm-1, consistent with bond-bending Ge–H2 from nearest neighbor Ge atoms at the edges of the crystalline sheets are observed.6a We do not observe any broad intense, Ge–O–Ge or Ge–O vibrational modes that should occur between 800 and 1000 cm-1. Furthermore, there is a consistent red shift in the Ge–H stretching frequency and the major Ge–H wagging mode upon increasing Sn incorporation (Fig. S4). For example, at a maximum of 9% Sn, the Ge–H stretching frequency decreases from 2002 cm-1 to 1988 cm-1 and the major Ge–H wagging mode decreases from 482 cm-1 to 470 cm-1. This is consistent with previously reported shifts observed in alloyed Sn/Ge:H thin films.19 Here, it has been established that the lower electronegativity of neighboring Sn atoms as compared to Ge reduces the Ge-H vibrational frequency. Taken together, after deintercalation in HCl, the Sn atoms are terminated with –OH groups, and the Ge atoms remain terminated with –H. Raman spectroscopy further confirms the alloy formation of the 2D network. Previous measurements showed that GeH exhibits E2 Ge-Ge Raman active modes at 302 cm-1, and an A1 out-of-plane Raman active mode at 228 cm-1, which were assigned by the close agreement with ab initio calculations.6a With increasing tin, there is an increasing shift in both Raman modes to lower energies, and at 9% tin, the E2, and A1 mode shift to 299 cm-1 and 221 cm-1, respectively (Fig. 2b). Furthermore, the FWHM of the E2 vibration also increases with increasing tin concentration (Fig. S5). This broadening in the FWHM and the decrease in vibrational phonon energy with increasing tin are also observed in 3D tin/germanium alloys, which occur due to the increasing average Ge-(Ge/Sn) bond length, as well as the increasing disorder in the effective mass distribution, respectively.19-20

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 13

was performed by fitting the Kubelka-Munk absorbance assuming direct-allowed, direct-forbidden, indirectallowed and indirect-forbidden gaps using the Tauc/Davis-Mott models of 2D and 3D densities of states (Fig. S6-8).22 For the 2D DOS model, a direct allowed band gap is characterized by a discontinuity in the absorbance resulting from the step function of the absorbance A(ħω) at photon energy ħω. On the other hand, an indirect allowed band gap is characterized by an absorbance function proportional to ħω – Eg’ ± Ep where Eg’ is the energy of the indirect band gap and Ep is energy of the corresponding phonon mode. However, it has been experimentally established that Tauc/Davis Mott approximations very often do not definitely determine the direct or indirect nature for materials with 2D densities of states.22 The determination of a band gap character of these materials (direct/indirect) is further complicated in these materials by the presence of defects or disorder in the lattice, which is evident from the broad Urbach edge at the absorption tail.

Figure 3. (a) Diffuse reflectance absorption spectrum of Ge1-xSnxH1-x(OH)x plotted in terms of the Kubelka-Munk function, F(R), and the photon energy. (b) Optical band gap of Ge1-xSnxH1-x(OH)x versus the Sn concentration.

Figure 2. (a) Transmission mode FTIR spectra of Ge1xSnxH1-x(OH)x (x=0-0.09). (b) Raman spectrum of Ge1xSnxH1-x(OH)x with the A1 mode highlighted as an inset. Raman shifts of the (c) A1 and (d) E2 phonon modes of Ge1xSnxH1-x(OH)x as a function of Sn concentration.

To determine the influence of Sn-OH substitutions on the optical band gap of GeH, we performed diffuse reflectance absorption measurements (DRA) on powders (Fig. 3a). The silver-black materials have a broad absorption over all visible wavelengths, and there is a systematic decrease in the absorption band edge with increasing tin concentration (Fig. 3b, S9). The linear fitting of the absorption edges of these materials results to a general decrease in the band edge from 1.59 eV for GeH down to the following band gap energies at different Sn concentrations: 1.52 ± 0.01 eV at 4%, 1.45 ± 0.01 eV at 7% and 1.38 ± 0.02 eV at 9% Sn. This shift in band gap shows a similar trend to 3D germanium/tin alloys where a decrease in band gap with increasing tin percentage is also observed.21 An attempt to determine whether these materials are direct band gap materials

The air stability of these alloys was probed via a 15day air exposure of a representative Ge0.91Sn0.09H0.91(OH)0.09 sample. After exposure to air, the absorption onset of the material decreases by ~0.2 eV (Fig. 4a). This occurs in both diffuse reflectance (Fig. 4a) and transmission (Fig. S10) experiments, suggesting that the sample is changing uniformly. A sharp vibrational feature around 847 cm-1 and a broad vibrational feature around 900 cm-1 also emerged in the FTIR spectrum (Fig. 4b). These vibrational modes correspond to Ge-O vibrations, that have been observed in amorphous GeOx.23 No other changes were observed in the FTIR spectrum (Fig. S11). X-ray photoelectron spectroscopy (XPS) (Fig. 4c) also confirmed the oxidation of germanium through the appearance of a shoulder centered around 1219 eV which is indicative of the presence of Ge2+/3+ 2p3/2 peak. A new shoulder centered around 487 eV, corresponding to a Sn4+ 3d5/2 peak, was observed after 15 days of exposure which indicates that tin on the surface is also oxidized (Fig. 4d). Since no new peaks were observed in the XRD after exposure to air, this suggests that the flakes are partially transforming to an amorphous GeOx/SnOx phase.

ACS Paragon Plus Environment

Page 5 of 13

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 4. (a) Diffuse reflectance absorption measurements, (b) FTIR spectra, and XPS spectra of the (c) Ge 2p3/2 and (d) Sn 3d5/2 of Ge0.91Sn0.09H0.91(OH)0.09 before (red) and after 15 day exposure to air (black) and after subsequent 1 M HCl washing (green).

This amorphous GeOx/SnOx layer can be readily removed via an HCl rinse. We rinsed these 15-day samples in 1 M HCl for 30 minutes, washed with acetonitrile and dried under vacuum. After washing, the absorption onset reverted back to 1.38 eV (Fig. 4a) and the vibrational modes at 847 cm-1 and 900 cm-1 corresponding to GeOx disappeared in the FTIR spectrum (Fig. 4b). Additionally, the XPS spectrum only showed Ge1+ 2p3/2 peak (Fig. 4c) and majority of the intense Sn4+ 3d5/2 peak also disappeared (Fig. 4d). Previous studies have established that GeOx can be washed with HCl forming a soluble GeCl4 species, creating a Ge vacancy and terminating the neighboring Ge atoms with hydrogen.23 While there have been considerably fewer studies with tin, these results show that a similar surface cleaning mechanism likely occurs with the removal of amorphous SnOx.

Figure 5. (a, left) Image of the actual device based on Ge0.91Sn0.09H0.91(OH)0.09 and (a, right) schematic illustration and dimensions of the photodetector (not drawn to scale). (b) Typical I-V plot of the Ge0.91Sn0.09H0.91(OH)0.09 photodetector in dark and at 600 nm illumination. (c) Photocurrent response at selected illumination wavelengths and (d) wavelength-dependent normalized photocurrent for the Ge0.91Sn0.09H0.91(OH)0.09 photodetector compared to GeH at a 3V bias voltage.

As has been established in 3D germanium/tin alloys, photoconductivity measurements are a powerful method to probe the effect of tin substitution on the optoelectronic properties of semiconductor materials.11b,13 We performed two-probe photoconductivity measurements on single crystals of GeH and Ge0.91Sn0.09H0.91(OH)0.09. 100 nm Ag/20 nm Au contacts were deposited via an e-beam evaporator using a 25 µm shadow mask (Fig 5a). These non-extrinsically doped crystals exhibit linear current vs. voltage characteristics, indicative of ohmic contacts, with resistances in the 10 GΩ range (Fig 5b). A significant photoconductive response is observed in the pure GeH device when exciting with above band gap light, and occurs rapidly with photoresponse rise times